20.01.2013 Views

THE ROLE OF THE SHAFT IN THE GOLF SWING

THE ROLE OF THE SHAFT IN THE GOLF SWING

THE ROLE OF THE SHAFT IN THE GOLF SWING

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

J. Bmmechanics Vol. 25, No. 9, pp. 975-983. 1992.<br />

Printed in Great Britain<br />

<strong>THE</strong> <strong>ROLE</strong> <strong>OF</strong> <strong>THE</strong> <strong>SHAFT</strong> <strong>IN</strong> <strong>THE</strong> <strong>GOLF</strong> SW<strong>IN</strong>G<br />

RONALD D. MILNE* and JOHN P. DAVIS?<br />

Departments of *Engineering Mathematics and JCivil Engineering, University of Bristol,<br />

Bristol BS8 ITR, U.K.<br />

CQZl-9290/92 %S.oO+.OO<br />

Pergamon Press Ltd<br />

Abstract--Current marketing of golf clubs places great emphasis on the importance of the correct choice of<br />

shaft in relation to the golfer. The design of shafts is based on a body of received wisdom for which there<br />

appears to be little in the way of hard evidence, either of a theoretical or experimental nature. In this paper<br />

the behaviour of the shaft in the golf swing is investigated using a suitable dynamic computer simulation<br />

and by making direct strain gauge measurements on the shaft during actual golf swings. The conclusion is,<br />

contrary to popular belief, that shaft bending flexibility plays a minor dynamic role in the golf swing and<br />

that the conventional tests associated with shaft specification are peculiarly inappropriate to the swing<br />

dynamics; other tests are proposed. A concomitant conclusion is that it should be difficult for the golfer to<br />

actually identify shaft flexibility. It is found that if golfers are asked to hit golf balls with sets of clubs having<br />

different shafts but identical swingweights the success rate in identifying the shaft is surprisingly low.<br />

<strong>IN</strong>TRODUCTION<br />

‘We challenge anyone to prove the shaft is not the<br />

most important component of a golf club’. This bold<br />

statement accompanied a recent advertisement for<br />

golf clubs in a popular golf magazine and is typical of<br />

the emphasis that is currently being placed on the role<br />

of the shaft in golf club design. Shafts are currently<br />

being manufactured to closer tolerances than ever<br />

before and in a wider range. What is in doubt is the<br />

basis of the design philosophy. Roughly speaking, the<br />

received wisdom dictates that the more flexible shaft is<br />

better suited to the weaker player, while the profes-<br />

sional or good amateur should use a stiffer shaft.<br />

Considered as an empirical result arising from the<br />

long development of golf, it must be taken seriously;<br />

however, within the context of the dynamics of the<br />

golf swing, it is not an obvious conclusion. A typical<br />

explanation (Tolhurst, 1989) of the effect of shaft<br />

flexibility runs along the following lines (for those<br />

unfamiliar with golf, a glossary of terms is given in<br />

Appendix 1). The shaft is initially bent backwards in<br />

the early part of the downswing and then recovers just<br />

before impact. If the shaft is too flexible for the golfer,<br />

it springs forward too far, thereby increasing the<br />

effective loft and closing the clubface; if too stiff, it is<br />

still bent backwards, the loft is decreased and the face<br />

is open. In this sense the shaft should be matched to<br />

the golfer’s swing speed and hand action. Claims are<br />

sometimes made for increased clubhead speed at im-<br />

pact, assuming that the golfer’s timing can utilise the<br />

springing back of the shaft. The link between change<br />

in loft and closing or opening of the clubface is due to<br />

the clubhead being attached to the shaft at the lie<br />

angle (Appendix 1). The actual extent to which the loft<br />

is changed by shaft bending can be controlled by the<br />

location of the so-called kick-point which relates to<br />

the shape of the bent shaft at impact; a shaft with a<br />

Receiued in final form 8 January 1992.<br />

915<br />

low kick-point shows greater curvature near the tip<br />

region than one with a high kick-point.<br />

Shaft design is currently verified by three basic tests.<br />

In the first the shaft is clamped at its butt over a<br />

standard length and then loaded transversely by a<br />

weight near the tip; the tip deflection yields a measure<br />

of the shaft bending stiffness, and the location of the<br />

maximum deviation of a chord joining the butt and<br />

tip from the deflection curve yields the kick-point. The<br />

second test measures the tip rotation due to an<br />

applied tip torque, so yielding the torsional stiffness.<br />

The third test measures the fundamental transverse-<br />

vibration frequency of the clamped shaft carrying a tip<br />

mass equivalent to a clubhead.<br />

The purpose of this study was to gain an under-<br />

standing of the flexing behaviour of the shaft during<br />

the swing by constructing a suitable dynamic model<br />

and to test the validity of the model experimentally.<br />

Shaft flex may affect the feel of a golf club during the<br />

swing; this difficult area warrants a full investigation<br />

on its own but results of a rather limited study are<br />

reported here since they add some weight to the<br />

conclusions.<br />

Computer simulation<br />

METHODS<br />

Flexibility of the shaft is manifest in bending and<br />

twisting. The former is the subject of careful design, so<br />

that desired characteristics are obtained, whereas the<br />

latter is seen as a necessary evil, which must be<br />

minimised. In this study emphasis is on bending flex-<br />

ibility; consequently, the dynamic model is a develop-<br />

ment of the well-established plane two-link model of<br />

the golf swing (Budney and Bellow, 1979; Daish,<br />

1972). This consists of an arm link driven by a shoulder<br />

torque and a club link driven by a wrist torque; the new<br />

element added to the established model is the bending<br />

flexibility of the club link. Although the bodily mo-<br />

tions of a golfer are quite complicated, numerous


916 R. D. MILNE and J. P. DAVIS<br />

photographic and kinematic studies have shown that<br />

the downswing is, indeed, executed more or less in a<br />

plane. One aspect of the real three-dimensional swing,<br />

which is imported into the essentially two-dimen-<br />

sional or plane model, derives from the fact that the<br />

centre of mass of the clubhead does not lie on the<br />

shaft. Due to the large centrifugal forces involved in<br />

the swing, this has an important effect on shaft bend-<br />

ing and, consequently, for this reason, it is necessary<br />

to allow for rotation of the clubhead in the plane of<br />

the swing, although this rotation has no direct effect<br />

on the momentum balances or impact mechanics in<br />

the plane. A description of the full equations of mo-<br />

tion together with details of the computation are set<br />

out in Appendix 2.<br />

One integral feature of the model is crucial and<br />

needs emphasis. The shaft has a bending stiffness,<br />

which is measured in the standard cantilever bending<br />

test already referred to. However, during the swing,<br />

particularly near impact when the centrifugal forces<br />

are large, the shaft also acquires an additional bend-<br />

ing stiffness due to the fact that the shaft is under<br />

considerable tension. In a full drive or similar-<br />

distance shot this bending stiffness due to tension is, in<br />

the 40 ms or so before impact, large enough to be<br />

comparable to the conventional bending stiffness.<br />

Strain gauge measurements<br />

Three golfers, labelled A, B and C, with handicaps<br />

of 11, 5 and 2, respectively, were used in these<br />

experiments; golfers B and C showed a high degree of<br />

consistency in repeating their swings. The club used<br />

for the tests was a metal driver with R shaft and<br />

having 10.5” of loft. Groups of foil strain gauges<br />

(Showa, 2 mm) were bonded to the shaft at three<br />

stations to measure two bending moments at each<br />

station, parallel and normal to the clubface. Tension<br />

and torque were also measured at one station. The<br />

gauges were calibrated directly for these moments and<br />

forces by appropriate static loadings of the shaft. The<br />

strain gauge amplifiers (Fylde mini-bal and mini-amp<br />

systems) transmitted data via an analogue/digital con-<br />

vertor (Barr-Brown PCI-20089) directly onto a com-<br />

puter (Toshiba 3100SX) controlled by data acquisi-<br />

tion software (Labtech Notebook); data was stored in<br />

a form suitable for subsequent analysis using the<br />

spreadsheet Lotus l-2-3. The typical run extended<br />

over 10 s beginning at address and ending at follow-<br />

through. Sampling was at the rate of 200 readings per<br />

second; at this sampling rate it was not possible to<br />

read all the gauges simultaneously and successive<br />

swings were used. Graphical displays of data could be<br />

obtained on the Toshiba screen immediately after a<br />

run in order to check for failure or inconsistency.<br />

Initial analysis located the region of interest from just<br />

before top of swing to shortly after impact by refer-<br />

ence to the obvious point of impact with the ball. The<br />

swing was also observed by a single high-speed video<br />

system running at a rate of 200 frames per second and<br />

shutter speed of l/10,000 of a second placed facing the<br />

golfer; the video system was an NAC Model HSV400<br />

with associated X Y-coordinator Model V-78-E<br />

coupled to an IBM-compatible PC using the motion<br />

analysis package MOVIAS (version 3). A video re-<br />

cording was also made along the line-of-flight direc-<br />

tion to establish the angle of the golfer’s swing plane.<br />

The video recordings allowed the gauge data to be<br />

correlated with swing position through the common<br />

time of impact and also allowed the construction of<br />

stick diagrams for the swing and the estimation of<br />

clubhead velocity and acceleration. The most import-<br />

ant information obtained from the recordings gave<br />

the orientation of the clubhead relative to the plane of<br />

the swing. This was needed to transform the parallel-<br />

to-clubface and normal-to-clubface bending-moment<br />

components to in-swing-plane and out-of-swing<br />

plane components since only the in-swing-plane com-<br />

ponents are relevant to the simulation. This aspect of<br />

the experimental method was the least satisfactory.<br />

The clubface was marked and the video image was<br />

clear; nevertheless, an accurate measurement of the<br />

orientation was difficult. Typically, the clubhead<br />

rotates little during the downswing until within about<br />

50 ms of impact when it rotates through approx-<br />

imately a right-angle to square the clubface at impact.<br />

A simple smoothing routine was used to obtain both<br />

in-plane and out-of-plane bending moments. Out-of-<br />

plane bending moments were very small except near<br />

impact, where the values were consistent with out-of-<br />

plane bending of the shaft due to centre-of-mass offset:<br />

the behaviour of the resolved bending moments serves<br />

to give confidence in the plane-swing model. Were<br />

experiments of this type to be repeated, a more accur-<br />

ate and automatic way of measuring orientation,<br />

perhaps using laser marker techniques, would be<br />

advisable. Questions of cost precluded such an ap-<br />

proach in these experiments.<br />

RESULTS<br />

A set of results is shown in full for golfer B only;<br />

results for golfers A and C naturally differ in detail but<br />

not enough to alter the overall picture as far as the<br />

shaft behaviour is concerned. A typical set of results<br />

for a swing simulation is shown in Fig. 1. Clubhead<br />

speed at impact estimated from the video via the<br />

MOVIAS program was 41.4 ms-‘. The initiation<br />

time for the swing simulation was chosen as 400 ms<br />

before impact, at which point in the backswing the<br />

shaft is straight and the wrist torque is just beginning<br />

to reverse: angular velocities at initiation were estim-<br />

ated from the video recording. The diagrams have<br />

been extracted from a continuous screen display to<br />

illustrate the main features. The club used for the<br />

simulation is a driver; a long iron club would show a<br />

very similar swing except that the bending forward of<br />

the shaft before impact would be much reduced. The<br />

stick diagrams in Fig. 1 show the shaft deflection to<br />

scale, and it is difficult to perceive detail; so, the shaft


Time before impact 68 ms Time before impact 39ms<br />

Time before impilct l9ms<br />

Time after impact IOms Time after impact 24ms<br />

Role of shaft 911<br />

Sha ft.c+fkct+n<br />

wn’f’ed ’ bmes<br />

Driver-Iaft 10.5&9 lnyuct &city 41.6mIs Drier-loft Kl.5dcp Imct velocity Ll.Cm/s<br />

Clubface aqk ll.Bdq Ball vebcity 5X6m/s Ctubfact aqk 11.8de9 Ball velocity 5Z6m/s<br />

deflection is also shown magnified five times. These<br />

bending shapes are plotted in such a way that they<br />

have the property of being orthogonal (with respect to<br />

the club mass distribution) to a rigid-body rotation<br />

about the wrist hinge (Appendix 2). Other outputs<br />

from the simulation, shown in Fig. 2, are clubhead<br />

speed, clubhead deflection measured relative to a tan-<br />

gent at the butt end and hand force. The hand force is<br />

the internal vector force at the wrist hinge acting in<br />

opposite directions on the arm and club links; in<br />

Fig. 2 the components of this force are shown resolved<br />

along and normal to the instantaneous directions of<br />

both arm and club links. Figure 2 also shows the<br />

tension measured in the shaft at its mid-point: meas-<br />

ured mid-point torque is not shown, as it did not<br />

exceed 0.5 Nm until impact. In Fig. 3 are shown the<br />

simulated and measured in-swing-plane bending mo-<br />

ments at three stations along the shaft-at 10% (top),<br />

Fig. 1. Swing simulation showing shaft deflection.<br />

r-i!-<br />

1<br />

50% (middle) and 90% (bottom) of shaft length from<br />

the wrist hinge.<br />

Shoulder and wrist torques for the simulation ap-<br />

propriate to golfer B were synthesised from the strain<br />

gauge data in the following way. The shoulder torque<br />

can be estimated from the measured shaft tension by<br />

utilising the fact that the force exerted on the shaft by<br />

the hands is almost wholly axial to the shaft through-<br />

out the swing, the tangential component of the hand<br />

force being at most about 10% of the total; hence,<br />

shaft tension is little different from the total hand<br />

force, a result which can readily be demonstrated<br />

theoretically. By using a point in the swing at which<br />

the angular acceleration of the arm link is zero (about<br />

70 ms before impact) the inertia of the arms/torso is<br />

eliminated from the estimate. The agreement between<br />

axial (club) hand force and mid-shaft tension (Fig. 2) is<br />

very satisfactory when it is noted that the tension at


978 R. D. MILNE and J. P. DAVIS<br />

tim ms time ms<br />

- Axial<br />

---- Kangenttil<br />

‘-‘- shaft tension<br />

Fig. 2. Swing simulation outputs and measured mid-shaft<br />

tension.<br />

10. ______- ----__ ___<br />

--._<br />

g-1: 1 -f----- % t<br />

-10<br />

\_I--<br />

0 100 ZW 300 100 SW<br />

Time ins<br />

Fig. 3. Simulated and measured shaft bending moments.<br />

mid-shaft is approximately 90% of the tension at the<br />

top. The wrist torque can obviously be estimated from<br />

the measured bending moment at the top station; the<br />

nearer this station is to the wrist hinge the more<br />

accurate the estimate. Based on this evidence the<br />

shoulder torque was applied as a ramp function with<br />

rise time 110 ms and maximum amplitude 83 N m and<br />

the wrist torque was applied as a ramp function with<br />

rise time 275 ms and maximum amplitude 25 Nm.<br />

These simplified torques were not intended to match<br />

the golfer’s inputs exactly. The inputs for golfers A<br />

and C were qualitatively similar to those for golfer B<br />

and the level of agreement between simulation and<br />

measurement much the same; golfer C was a very long<br />

hitter generating a clubhead speed of 48 m s-l.<br />

<strong>IN</strong>TERPRETATION <strong>OF</strong> RESULTS<br />

For the purpose of describing the shaft behaviour<br />

the swing may be divided into three phases. In the first<br />

phase, which extends from about 230 ms before im-<br />

pact (top of the swing) to about 130 ms before impact,<br />

the application of the wrist torque bends the shaft<br />

backwards, the maximum deflection occurring near<br />

the end of this phase. In the second phase, which<br />

consists of the last 130 ms or so to impact, an unfold-<br />

ing of the system due to momentum transfer takes<br />

place. The shaft behaves in a quasi-steady manner,<br />

gradually straightening and then bending forward due<br />

to the anticlockwise centrifugal torque applied at the<br />

lower end of the shaft by the offset of the centre of<br />

mass of the clubhead behind the shaft centre-line. This<br />

torque is closely represented by the bending moment<br />

at the bottom station (Fig. 3); near impact the effective<br />

clubhead weight is some 150 times its actual weight,<br />

so the magnitude of the torque applied is some 3 N m<br />

for each centimetre of centre-of-mass offset. The main<br />

effect is to give the club an additional dynamic loft at<br />

impact with a corresponding closing of the clubface.<br />

The nett clubface angle at impact with the ball is<br />

recorded by the simulation (Fig. 1). It is important to<br />

note that the position of the centre of mass of the<br />

clubhead relative to the hands is along the local<br />

centrifugal vector, so that if the shaft is bent forward<br />

then the club butt-to-arm angle is correspondingly<br />

increased and vice versa (see also Fig. 4). In phase<br />

three the effect of impact on the shaft is to excite a few<br />

cycles of bending vibration at a frequency of about<br />

30 Hz. Some of the energy at impact is spent in<br />

exciting the shaft; for example, given the same set of<br />

input torques, the simulation shows that a flexible<br />

shaft produces an initial ball velocity which is about<br />

4% lower than that produced by an equivalent ideal-<br />

ised ‘rigid’ shaft.<br />

The computer simulation necessarily incorporates<br />

a model of the dynamic characteristic or impedance of<br />

the wrist joint, and this is an area where information is<br />

lacking: experimental work such as that carried out by<br />

Brown et al. (1982) on other human joints is badly<br />

needed. When gripped by a golfer, the club is very far<br />

from being clamped, as in the standard bending and<br />

vibration tests. The stiffness or the in-phase compon-<br />

ent of the wrist impedance torque is probably quite<br />

small when, near impact, the joint is relaxed and<br />

rapidly developing. The damping or the out-of-phase<br />

component is large enough near impact to actually<br />

apply a retarding torque. The measured top station<br />

bending moment in Fig. 3 clearly shows the decrease<br />

of the wrist torque approaching impact as the rota-<br />

tion speed of the joint rapidly increases. This behavi-<br />

our was observed for all the golfers tested and has<br />

been observed by others (Budney and Bellow, 1979);<br />

details of the impedance used in the model are given in<br />

Appendix 2. By varying the impedance-damping level<br />

and using the same demand inputs, the simulator can<br />

arrive at impact with wrist torques which vary from<br />

decelerating (the usual case.) through neutral to accel-<br />

erating. The results from three simulations of this<br />

type, based on the swing of golfer B but taken over a<br />

rather extreme range for emphasis, are shown in<br />

Fig. 4. Note that the clubface angle at impact is the


-_<br />

Role of shaft 979<br />

r<br />

SWI defkction<br />

1982; Cochran and Stobbs, 1969) and confirm an<br />

maaniti& 5 times<br />

lime betore impact 4 ms<br />

earlier description of shaft behaviour which was based<br />

lIecekr&g wrist torque<br />

solely on well-founded physical reasoning (Williams,<br />

at inplct I- mNml<br />

1969).<br />

!<br />

DISCUSSION<br />

The main aim of the work reported here was to<br />

elucidate, in a qualitative way, the dynamic flexing<br />

1, -5 0 cm<br />

behaviour of the golf shaft during the swing. In the<br />

Drim-loft m.5* Impct vrbcityQ.9m/s<br />

simulation the precise torques generated by the<br />

Clubface u?pk B8dcg Bali nkcity 566 m/s<br />

‘golfer’ are open to question but the inertial and<br />

he before itquct 3ms<br />

MC&a/ wrist ivrque<br />

at impact (0 Nm)<br />

1,<br />

imposed forces and torques applied to the shaft itself,<br />

L-r<br />

as verified by the strain gauge measurements, are<br />

undoubtedly typical of the golf swing and, hence, the<br />

level of confidence that may be placed in the simulator<br />

is high. The simulator could be used as a design tool<br />

to explore in detail the interaction between a range of<br />

shafts and ‘golfers’ as represented by their input<br />

torques. No attempt is made to do this here, but it is<br />

worth remarking that widely different shoulder<br />

-5 0 cm 5 torques lead to similar simulated swings provided the<br />

Driver-loft 10.5dq Impct v&city 122m/s<br />

same amount of work is done: this is not a surprising<br />

Ctubfaa an9k 12&q Ball nlocity 58.4 m/s<br />

result if one remembers that the swing is the outcome<br />

Time be/on impact 2ms<br />

of a double time integration and is reflected in the<br />

observation that, on the golf course, very different<br />

looking golf swings can produce similar results.<br />

The burden of the conclusions arising from the<br />

study is that shaft flexibility does not play an important<br />

dynamic role in the swing and the principal effect<br />

&her-btt M5de9 1-t bvtocity 43.5m/s<br />

of flexibility, namely, change in the effective loft at<br />

impact, could be estimated from static considerations.<br />

A simple test which would serve this purpose is illustrated<br />

in Fig. 5(a); this simulates the behaviour of the<br />

shaft of a wooden club just before impact when condi-<br />

Cl&face qle 13.Ldq BaX v&city 60 1 m/s<br />

tions are quasi-steady and the shaft is subject to a<br />

Fig. 4. Effect of wrist torque on shaft shape at impact. large axial load. The outcome of the test is a direct<br />

measure of the change in the effective loft (per unit<br />

centre-of-headmass offset) which this shaft would give<br />

result of two opposing effects, the change in butt-to-<br />

arm angle. being counteracted by the change in rota-<br />

tion at the shaft tip. The impact velocity is increased<br />

only by 2.6 m s- ‘, which is consistent with the fact<br />

that, for the whole swing to impact, approximately<br />

92% of the work done comes from the shoulder<br />

torque (via the torso and legs) and only 8% from the<br />

wrist torque.<br />

The clamped bending frequency of a driver is typi-<br />

cally of the order of 4.5 Hz, whereas the bending<br />

frequency observed for a club freely hinged near its<br />

butt end as illustrated in Fig. 5(b) is of the order of<br />

25 Hz. The bending frequency for a club, freely rota-<br />

ting about a hinge near its butt end and having a<br />

speed of rotation which gives the head a speed of<br />

40 m s- I, is about 32 Hz as observed in the simulation<br />

after impact; in this situation the frequency is only<br />

weakly dependent on the shaft bending stiffness. The<br />

three phases described above are all observable in<br />

high-speed photographs of golfers’ swings (Maltby,<br />

(b) m<br />

Changeotlottdwtooftaotofcmtmof<br />

- cd clubhud lrom sh#t cantm IIM<br />

ShCftICplVOtCdCtbldtCndlOCdWIC<br />

rppScd vlc rigId bm&ct ct lip<br />

Appllcalon 01 wlcl torque un bc<br />

*mummd by applybIg mothnr load via<br />

� rtgld bmcket atUch#d at butt<br />

Lcwca ncxuml Vlbmtlon mode 01 Divomd<br />

ShM with bndrmu<br />

Fig. 5. Proposed shaft tests.


980<br />

and is related to the player’s clubhead speed at impact<br />

if the load W is given a value appropriate to the<br />

centrifugal force on the head; it automatically takes<br />

account of the change in club butt-to-arm angle al-<br />

ready referred to earlier. Application of a torque at the<br />

butt end can simulate the golfer’s wrist torque at<br />

impact. This test is reflected in the common practice<br />

amongst golfers of resting the toe of a club on the<br />

ground and then pushing the butt firmly downwards!<br />

Clearly, another test which should be included in<br />

shaft testing is the measurement of the fundamental<br />

bending-vibration frequency of a shaft-plus-head sus-<br />

pended freely from a pivot as illustrated in Fig. 5(b).<br />

To demonstrate the effect of an increase in the bend-<br />

ing stiffness due to centrifugal force, approximate<br />

vibration frequencies are shown for a rotating shaft-<br />

plus-head and for a rotating chain-plus-head, the<br />

chain having the same mass distribution as the shaft<br />

but, of course, zero conventional bending stiffness.<br />

If shaft flexibility is not dynamically significant then<br />

why is it apparently so important to the golfer? To<br />

begin to answer this question in a very preliminary<br />

way a series of trials was conducted with a small<br />

group of golfers consisting of four amateurs, with<br />

handicaps ranging from 13 to 5, and four profes-<br />

sionals. Three sets of clubs were specially prepared,<br />

each set having identical swing weights but fitted with<br />

R, S and X shafts, two sets of five-irons (with solid<br />

back and peripherally weighted heads) and a set of<br />

drivers. The decals on the shafts were covered by a<br />

code letter and the participants were instructed not to<br />

bend the shaft or to waggle the club but simply to hit<br />

golf balls. The results were somewhat surprising. The<br />

amateurs were unable to distinguish one shaft from<br />

another either for woods or irons; typically, if pressed<br />

as to which shafts they preferred, they would chose the<br />

R and X over the S or, if asked for an order of<br />

flexibility, would give an almost random selection.<br />

Striking of the ball did not seem to be greatly affected,<br />

although no measurements were made of ball traject-<br />

ory. The professionals were not much better at dis-<br />

criminating between the irons but were quite success-<br />

ful with the woods. However, even here success was<br />

not complete, although the R shaft could generally be<br />

distinguished from the S and X shafts. These very<br />

simple and restricted tests do suggest that more exten-<br />

sive and significant work should be undertaken in this<br />

area. Recently, Pelz (1990) has conducted ‘blind’ golfer<br />

trials to compare the performance of steel and graph-<br />

ite shafts of differing flexibilities and Van Gheluwe et<br />

al. (1990) have shown that the kinematics of a golfer’s<br />

swing appears to be independent of the shaft material.<br />

CONCLUSIONS<br />

Do the results of this study bear out the accepted<br />

explanation of the effect of shaft flexibility given in the<br />

Introduction? The study confirms that flexibility will<br />

produce a change of loft (and close clubface) at im-<br />

pact but shows that these are really quasi-static effects<br />

R. D. MILNE and J. P. DAVIS<br />

and not dependent on a dynamic ‘springing back’ of<br />

the shaft from its bent position at the start of the<br />

swing. It is difficult to give any credence to the sugges-<br />

tion that the shaft can actually be bent backwards at<br />

impact and no photographic evidence exists to sup-<br />

port this contention. Kick-point should be replaced<br />

by a measure of shaft-tip rotation due to an offset<br />

axial load. Nor does springing back or ‘whipping’ of<br />

the shaft play a dynamic role in the all-important pre-<br />

impact area, although the ‘feel’ of the violent flexing of<br />

the shaft after impact may give the golfer the impres-<br />

sion that it does. Indeed, clubhead speed at impact is<br />

virtually unchanged if, using the simulator, the shaft is<br />

replaced by a heavy chain about 60 ms before impact.<br />

Perhaps the major source of misunderstanding about<br />

the role of the shaft is associated with the idea that the<br />

club is ‘gripped’ by the golfer, resembling the condi-<br />

tion of the built-in butt end of the cantilever test,<br />

combined with a lack of appreciation of the magni-<br />

tude and speed of response of any forces and moments<br />

applied by the player, which could significantly affect<br />

the course of the swing in the pre-impact area.<br />

The implications of this study for club design are<br />

somewhat problematical. It does appear that it may<br />

not be necessary to provide a range of shafts for iron<br />

clubs and possibly not even for wooden clubs. It is<br />

possible to argue that the ‘feel’ imparted to a club<br />

through shaft flexibility may actually affect the torque<br />

inputs of the player but the golfer tests gave no<br />

indication of this: there is room here for further study.<br />

If one takes the view that shaft flexibility is not of<br />

importance dynamically, it then represents an un-<br />

wanted swing variable and one is led to adopt a stiff<br />

shaft, a conclusion also reached by Pelz (1990). The<br />

effective loft that a golfer may have enjoyed with a<br />

fiexible shaft could then be reinstated (without the<br />

associated change in clubface angle) by an appropri-<br />

ate adjustment of nominal club loft. Do current stiff<br />

shafts represent an upper limit or could stiffer shafts<br />

be used? If the latter, then it would be possible to use a<br />

light, thin-walled shaft tapering to about twice the tip<br />

diameter of the currently available shafts coupled with<br />

a return to hosel-in-shaft fabrication; such a shaft<br />

would have considerably higher torsional stiffness<br />

than the currently available shafts.<br />

Finally, the point should be made that the fixity<br />

conditions of a shaft held in a golf machine are very<br />

different from those for a golfer and, therefore, the<br />

shaft performance data derived from this source do<br />

not necessarily apply to the ‘real’ golf swing.<br />

Acknowledgements-The authors would like to express their<br />

gratitude to Dunlop Slazenger International Ltd for finan-<br />

cial aid and for supplying golf clubs.<br />

REFERENCES<br />

Brown, T. I, H., Rack, P. M. H. and Ross, H. F. (1982) Forces<br />

generated at the thumb interphalangeal joint during im-<br />

posed sinusoidal movements. .I. Physici. 332, 69-85.<br />

Budney, D. R. and Bellow, D. G. (1979) Kinetic analysis of a<br />

golf swing. Res. Quart. Exercise Sport 50, 171-179.


Cochran, A. and Stobbs, J. (1969) The Search for the Perfect<br />

Swing. Heinemann, London, England.<br />

Daish, C. B. (1972) The Physics of Ball Games-Part II.<br />

English Universities Press, London, England.<br />

Maltby, R. (1982) Golf Club Design, Fitting, Alteration and<br />

Repair. Ralph Maltby Enterprises, Newark, NJ.<br />

Pelz, D. (1990) A simple scientific shaft test: steel versus<br />

graphite. In Proceedings of the First World Scientific Con-<br />

gress of Golf (Edited by Cochran, A. J.), pp. 264269.<br />

Chapman & Hall, London, England.<br />

Tolhurst, D. (1989) What you need to know before you buy.<br />

Golf Monthly Equipment Supplement, April 1989 issue.<br />

Van Gheluwe, B., Deporte, K. and Ballegeer, K. (1990) The<br />

influence of the use of graphite shafts on golf performance<br />

and swing kinematics. In Proceedings of the First World<br />

Scientific Congress of Golf (Edited by Cochran, A. J.), pp.<br />

258-263. Chapman & Hill, London, England.<br />

Williams, D. (1969) The Science of the Golf Swing. Pelham<br />

Books, London, England.<br />

APPENDIX 1<br />

GLOSSARY <strong>OF</strong> <strong>GOLF</strong><strong>IN</strong>G TERMS<br />

The golf club consists of three parts: the head, the shaft and<br />

the grip. The head is either a fairly narrow metal blade (iron)<br />

or a more extended, bulbous shape with a flattish front face<br />

(wood) made traditionally of wood but now also made in<br />

metal. The loft of the club is the angle which the front face of<br />

the head (clubface) makes with the vertical. The shaft is<br />

attached to the head by the hose1 or neck, which is a cylindri-<br />

cal extension of the head situated near one end of the<br />

clubface. The hose1 is angled to the head in such a way that<br />

when the base of the head is resting flat on the ground the<br />

shaft is at an angle to the horizontal called the lie angle. The<br />

centre of mass of an iron club is situated about halfway along<br />

the blade while the centre of mass of a wooden club is offset<br />

from the neck roughly along a line at 45” to the clubface. The<br />

shaft is either a steel tube or a solid composite tapering down<br />

from the grip end (butt) to the tip. The distribution of the<br />

taper controls the bending and torsional properties of the<br />

shaft. Shafts are commonly manufactured in three main<br />

categories R (regular), S (stiff) and X (extra stiff): within each<br />

category different weights and kick-point positions are<br />

offered. Typically, the ratio of clamped transverse vibration<br />

frequencies for R, S and X shafts is 1.00: 1.04: 1.08. The grip is<br />

simply a rubber tube which is glued to the butt of the shaft.<br />

At impact the clubface is said to be open ifit is pointing to<br />

the right of the target line and closed if it is pointing to the<br />

left: the efictioe lof is the actual angle of the clubface to the<br />

vertical at impact as opposed to the nominal loft of the club.<br />

Bending of the shaft in the swing plane near impact not only<br />

changes the loft but also opens or closes the clubface; for<br />

example, for a driver with lie angle of 55” the clubface will<br />

open (close) by an angle which is 70% of the change in the<br />

loft.<br />

The swing plane is the imaginary plane in which the player<br />

swings the clubhead and will generally be inclined somewhat<br />

steeper than the lie angle of the club.<br />

Due to the arrangement of bones and joints in the lower<br />

arms and wrists, the golfer is forced to roll or rotate his<br />

hands during the golf swing if he is to allow the wrist hinge to<br />

move freely: the face of the club lies roughly parallel to the<br />

swing plane at the top of the swing and is then rotated so as<br />

to be at right-angles to the swing plane at impact.<br />

APPENDIX 2<br />

<strong>THE</strong> MA<strong>THE</strong>MATICAL MODEL<br />

Equations of motion<br />

Figure Al illustrates the applied torques and rotation<br />

angles for the dynamic model. The relevant equations of<br />

Role of shaft 981<br />

Fig. Al. Applied torques and rotation<br />

dynamic model.<br />

angles for the<br />

motion, listed in the following, consist of<br />

(a) equations (1) and (2) representing the rate of change of<br />

angular momentum (including terms due to shaft flexibility)<br />

about the upper (fixed) pivot and about the wrist hinge,<br />

respectively, and<br />

(b) a variational equation (principle of virtual workj for<br />

shaft bending based on the small-deflection theory for slen-<br />

der beams including the effect on bending of tension in the<br />

shaft due to centrifugal forces. This integro-differential equa-<br />

tion is replaced by a number of ordinary differential equa-<br />

tions in time by representing the shaft bending displacement<br />

as a series of shape functions with time-dependent ampli-<br />

tudes (the Galerkin method) leading to equations (3), ,<br />

(n+2).<br />

Equation (1):<br />

(I,+m,l~)ti+ m,l,7,cos(+B)<br />

(<br />

-(i, EiCi+C)sinCO-B))$<br />

+ i E,cos(~-0)-<br />

i=1<br />

-(m.l,,csin(4-8)+<br />

-2 i E,&sin(4-0)$<br />

i=1<br />

+gsin0(l,m,+m,L)+ T,- T,=O,<br />

Equation (2):<br />

m,l,l;cos(#-0)-<br />

+(~, E,c,+c)cos))+T,rO.<br />

Equations (Z+i), i=l, . . . , n<br />

E,(sin(&e@+cos(4-e)i)+ i k,(T,+&)<br />

,=1<br />

+(cos(4-0)@-sin(+e)l) i; G,,rj+H,C<br />

j=l


982 R. D. MILNE and J. P. DAVIS<br />

- T,s;(o)/l,=o,<br />

G,,=f s ’ Ca(u)+mbls;(u)s;(u)du,<br />

+y,<br />

s 0<br />

* i<br />

s<br />

M,,=Pok P,(~W~,W du+wi(lb,W,<br />

0<br />

s<br />

1<br />

K,‘= [rk,(u)+m&;(u)s;(u)du, c=l,m,E,<br />

0<br />

ei(u)s;(u)s;(u)du,<br />

T,=Tt+imp@-4, e-4);<br />

dot denotes differentiation with respect to time, t,<br />

whereas prime denotes differentiation with respect to a,<br />

Z( 4) = distance of c of m of clubhead from shaft in<br />

swing plane,<br />

ei(u) = bending stiffness ratio of shaft, ei(0) = 1,<br />

EI = reference bending stiffness of shaft (at u = 0),<br />

&j=gcos(swp)<br />

(J = gravitational acceleration,<br />

{ swp = angle of swing plane to vertical,<br />

Jr = moment of inertia of arms/torso about upper<br />

pivot,<br />

I, = moment of inertia of club about wrist pivot,<br />

imp( ) = wrist impedance function,<br />

I, =length of arm link,<br />

1, = length of club link,<br />

L = distance of c of m of arm link from upper pivot,<br />

t = distance of c of m of club link from wrist pivot,<br />

m, = mass of club,<br />

m, =mass of clubhead,<br />

a = number of shape junctions,<br />

s,(u) = ith shape function,<br />

t = time,<br />

T, = torque at upper pivot (‘shoulder torque’),<br />

T, = torque at wrist pivot (%rist torque’),<br />

T”, =‘demand wrist torque’,<br />

a = material,damping constant for shaft,<br />

g(t)=angle of arm link to vertical,<br />

C(t) = amplitude of ith shape function,<br />

p,(u) =mass density ratio of shaft per unit length,<br />

p(O)= 1,<br />

p. = reference mass density of shaft (at u = 0),<br />

u = dimensionless distance along shaft from wrist<br />

pivot,<br />

$(t) = angle of club link axis to vertical (Fig. Al).<br />

First-orderform of equations of motion<br />

where<br />

f,=z,<br />

“+Z<br />

,TI m,‘({x~})i’=Ft({x~},Izt))<br />

)<br />

i,k=l, . . ..n+2.<br />

{xsl=(@, 4, L 52,. . . . 9 5.1.<br />

In this form the equations of motion are integrated forward<br />

in time using a fourth-order Runge-Kutta scheme.<br />

Momentum balance equations at impact<br />

During forward integration of the equations of motion,<br />

impact with the ball is detected by an interpolation of the<br />

clubhead path between time steps: the program enters a<br />

routine which solves the following algebraic equations representing<br />

conservation of momentum of the whole system<br />

and then reenters the integration program with a new set of<br />

initial values, the first time step being adjusted to lock onto<br />

the pre-impact time markers;<br />

“C7.<br />

C m,,({x~“‘})(zjf’-zy’)=m,uu,,<br />

‘==I<br />

i,k=l,...., n+2,<br />

n+2<br />

C u,(z~f)+ez~i))=u i<br />

i=1<br />

where<br />

{Ui} =(l,cosf+‘, l,cosfjP~), s,(l)cosf$“m),<br />

..., s ” (1)cosQ;““‘)<br />

$m) = (j+m) + y, $im) = &0 + ?,<br />

m, = mass of golf ball,<br />

v = initial ball velocity,<br />

e = coefficient of restitution,<br />

y =initial angle of ball flight path above horizontal,<br />

(im)*(i)s(‘)denote impact, initial and 6nal values.<br />

Shape functions-pre-processing program<br />

The n shape functions are defined by sr(u)=c;L: +a’,<br />

where the coefficients c,, are chosen such that, for all<br />

i,j= 1, . ..,n,<br />

1<br />

1<br />

~(~)s,(~)~d~+~~,(l)=4<br />

s 0<br />

p(u)si(u)s’(u)du+mssi(l)s,(l)=O, i#j.<br />

I 0<br />

That is to say, with respect to the mass distribution of the<br />

club, the shape functions are mutually orthogonal and are<br />

also orthogonal to the rigid-body rotation u. As a result, the<br />

left-hand side matrix cm,,] has non-zero entries on the<br />

diagonal and in the first row and first column only. Given<br />

the club mass and bending stiffness distributions, a preprocessing<br />

program generates the shape functions recursively,<br />

then computes the stiffness matrix [k,,]. The program<br />

also calculates the fundamental pivoted vibration frequency<br />

of the club [Fig 5(b)]. It is found that three shape functions<br />

are sufficient to give good accuracy for the frequency calcu-


Role of shaft 983<br />

lation and to deal with the changes in shape of the shaft for the three golfers, force demand being entirely embodied<br />

which occur throughout the swing simulation. in 7’:. An alternative model using a kinematic demand on<br />

club/arm angle and embodying also a non-linear stiffness<br />

Wrist impedance function term weighted towards the limit of wrist rotation gives very<br />

The hypothetical impedance function consists solely of the<br />

damping term const. (O-c$)‘, with the same constant used<br />

similar results in the simulation.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!